Fluid thread breakup

Last updated

Fluid thread breakup is the process by which a single mass of fluid breaks into several smaller fluid masses. The process is characterized by the elongation of the fluid mass forming thin, thread-like regions between larger nodules of fluid. The thread-like regions continue to thin until they break, forming individual droplets of fluid.

Contents

Thread breakup occurs where two fluids or a fluid in a vacuum form a free surface with surface energy. If more surface area is present than the minimum required to contain the volume of fluid, the system has an excess of surface energy. A system not at the minimum energy state will attempt to rearrange so as to move toward the lower energy state, leading to the breakup of the fluid into smaller masses to minimize the system surface energy by reducing the surface area. The exact outcome of the thread breakup process is dependent on the surface tension, viscosity, density, and diameter of the thread undergoing breakup.

History

The examination of droplet formation has a long history, first traceable to the work of Leonardo da Vinci who wrote: [1]

"How water has tenacity in itself and cohesion between its particles. […] This is seen in the process of a drop becoming detached from the remainder, this remainder being stretched out as far as it can through the weight of the drop which is extending it; and after the drop has been severed from this mass the mass returns upwards with a movement contrary to the nature of heavy things."

He thus correctly attributed the fall of droplets to gravity and the mechanism which drives thread breakup to the cohesion of water molecules.

The first correct analysis of fluid thread breakup was determined qualitatively by Thomas Young and mathematically by Pierre-Simon Laplace between 1804 and 1805. [2] [3] They correctly attributed the driver of thread breakup to surface tension properties. Moreover, they also deduced the importance of mean curvature in the creation of excess pressure in the fluid thread. Through their analysis, they showed that surface tension can behave in two ways: an elastic mechanism that can support a hanging droplet and a pressure mechanism due to capillary pressure that promotes thread breakup.

In the 1820s, the Italian physicist and hydraulic engineer Giorgio Bidone studied the deformation of jets of water issuing from orifices of various shapes. [4] Félix Savart followed in 1833 with experimental work, utilizing the stroboscopic technique to quantitatively measure thread breakup. [5] He noted that breakup is a spontaneous process, occurring without an external stimuli. This work allowed him to determine that droplets are produced from a jet flowing from a tank at a distinct rate inversely proportional to the nozzle radius and proportional to pressure in the tank. These observations facilitated Joseph Plateau's work that established the relationship between jet breakup and surface energy. [6] Plateau was able to determine the most unstable disturbance wavelength on the fluid thread, which was later revised by Lord Rayleigh to account for jet dynamics.

As the surface disturbance becomes large, non-linear theory must be applied. The behavior of jets with large disturbances was examined experimentally by Magnus and Lenard. [7] [8] Their experiments helped to characterize satellite droplets, droplets that are produced in addition to the large main droplet, through the introduction of high speed photography. High speed photography is now the standard method for experimentally analyzing thread breakup.

With the advent of greater computational power, numerical simulations have begun to replace experimental efforts as the chief means of understanding fluid breakup. However, difficulty remains in accurately tracking the free surface of many liquids due to its complex behavior. The most success has occurred with fluids of low and high viscosity where the boundary integral method can be employed as the Green's function for both cases is known. Dommermuth and Yue characterized irrotational, inviscid flow by this method as did Schulkes. [9] [10] Youngren and Acrivos considered the behavior of a bubble in a high viscosity liquid. [11] Stone and Leal expanded this initial work to consider the dynamics of individual drops. [12] For fluids of middling viscosity, full simulations using the Navier-Stokes equations are required with methods determining the free surface such as level-set and volume of fluid. The earliest work with full Navier-Stokes simulations was done by Fromm which focused on inkjet technology. [13] Such simulations remain an active area of research.

Physical mechanism of thread breakup

The process undergone by a fluid thread or jet undergoing breakup from a larger mass to a smaller mass. CombinedFluidThreadBreakupCylinders.svg
The process undergone by a fluid thread or jet undergoing breakup from a larger mass to a smaller mass.

The breakup process in a fluid thread or jet begins with the development of small perturbations on the free surface of the fluid. This is known as the linear theory of fluid thread breakup. These perturbations are always present and can be generated by numerous sources including vibrations of the fluid container or non-uniformity in the shear stress on the free surface. In general, these disturbances take an arbitrary form and are thus difficult to consider rigorously. It is therefore helpful to take a Fourier transform of the disturbances to decompose the arbitrary disturbances into perturbations of various single wavelengths on the surface of the thread. In doing so, this allows one to determine which wavelengths of the disturbance will grow and which will decay in time. [14]

The growth and decay of wavelengths can be determined by examining the change in pressure a perturbation wavelength imposes on the interior of the fluid thread. Changes to the internal pressure of the thread are induced by capillary pressure as the free surface of the thread deforms. Capillary pressure is a function of the mean curvature of the interface at a given location at the surface, meaning the pressure is dependent on the two radii of curvature that give the shape of the surface. Within the thinned area of a fluid thread undergoing breakup, the first radius of curvature is smaller than the radius of curvature in the thickened area, leading to a pressure gradient that would tend to force liquid from the thinned to thickened areas. However, the second radius of curvature remains important to the breakup process. For some perturbation wavelengths, the effect of the second radius of curvature can overcome the pressure effect of the first radius of curvature, inducing a larger pressure in the thickened regions than the thinned regions. This would push fluid back toward the thinned regions and tend to return the thread to its original, undisturbed shape. However, for other perturbation wavelengths, the capillary pressure induced by the second radius of curvature will reinforce that of the first radius of curvature. This will drive fluid from the thinned to the thickened regions and further promote thread breakup.

Radii of curvature in a thread undergoing the breakup process. Blue represents the first radius of curvature and red the second radius of curvature at the thinned and thickened locations. RadiiOfCurvatureFluidThreadBreakup.svg
Radii of curvature in a thread undergoing the breakup process. Blue represents the first radius of curvature and red the second radius of curvature at the thinned and thickened locations.

The wavelength of the perturbation is therefore the critical parameter in determining whether a given fluid thread will breakup into smaller masses of fluid. Rigorous mathematical examination of the perturbation wavelengths can lead to a relation showing which wavelengths are stable for a given thread as well as which perturbation wavelengths will grow most rapidly. The size of the fluid masses resulting from the breakup of a fluid thread can be approximated by the wavelengths of the perturbation that grow most rapidly.

Non-linear behavior

While linear theory is useful in considering the growth of small disturbances on the free surface, when the disturbances grow to have a significant amplitude, non-linear effects begin to dominate breakup behavior. The non-linear behavior of the thread governs its final breakup and ultimately determines the final shape and number of the resulting fluid masses.

Nonlinearity is captured through the use of self-similarity. Self-similarity assumes that the behavior of the fluid thread as the radius approaches zero is the same as the behavior of the fluid thread when it has some finite radius. Detailed understanding of non-linear thread behavior requires the use of asymptotic expansions to generate the appropriate scaling behavior. Numerous solutions have been found for the non-linear behavior of fluid threads based on the forces that are relevant in particular circumstances. [15] [16] [17]

Important parameters

How a fluid thread or jet undergoes breakup is governed by several parameters among which are the Reynolds number, the Weber number, Ohnesorge number, and the disturbance wavelength. While these numbers are common in fluid mechanics, the parameters selected as scales must be appropriate to thread breakup. The length scale most often selected is the radius of the fluid thread, while the velocity is most often taken to be the velocity of the bulk fluid motion. However, these scales can change based on the characteristics of the considered problem.

The Reynolds number is the ratio between inertia and viscous effects within the thread. For large Reynolds numbers, the effects of motion of the thread are much greater than viscous dissipation. Viscosity only has a minimal damping effect on the thread. For small Reynolds numbers, viscous dissipation is large and any disturbances are rapidly damped from the thread.

The Weber number is the ratio between inertia and surface tension effects within the thread. When the Weber number is large, the inertia of the thread is large which resists the tendency of surface tension to flatten bent surfaces. For small Weber numbers, the changes in the capillary pressure due to the surface disturbances is large and surface tension dominates thread behavior.

The Ohnesorge number is the ratio between viscous and surface tension effects within the thread. As it eliminates the effects of inertia and the need for a velocity scale, it is oftentimes more convenient to express scaling relationships in terms of the Ohnesorge number rather than the Reynolds and Weber number individually.

The perturbation wavelength is the characteristic length of the disturbance on the surface of the jet, assuming that any arbitrary disturbance can be decomposed via a Fourier transform into its constitutive components. The wavelength of the perturbation is critical in determining if a particular disturbance will grow or decay in time.

Special cases

Linear stability of inviscid liquids

The linear stability of low viscosity liquids was first derived by Plateau in 1873. [14] However, his solution has become known as the Rayleigh-Plateau instability due to the extension of the theory by Lord Rayleigh to include fluids with viscosity. Rayleigh-Plateau instability is often used as an introductory case to hydrodynamic stability as well as perturbation analysis.

Plateau considered the stability of a thread of fluid when only inertial and surface tension effects were present. By decomposing an arbitrary disturbance on the free surface into its constitutive harmonics/wavelengths, he was able to derive the a condition for the stability of the jet in terms of the perturbation:

where ω is the growth rate of the perturbation, σ is the surface tension of the fluids, k is the wavenumber of perturbation, ρ is the fluid density, a is the initial radius of the unperturbed fluid, and I is the modified Bessel function of the first kind. By computing the growth rate as a function of wavenumber, one can determine that the fastest growing disturbance wavelength occurs at:

The wavelength of maximum instability increases as the radius of the fluid thread increases. As importantly, unstable modes are only possible when:

Linear stability of viscous liquids

Reynolds and later Tomotika extended Plateau's work to consider the linear stability of viscous threads. Rayleigh solved for the stability of a viscous thread of viscosity without the presence of an external fluid. [18] Tomokita solved for the stability of a fluid thread in the presence of an external fluid with its own viscosity . [19] He considered three cases where the viscosity of the fluid thread was much greater than the external environment, the viscosity of the external environment was much greater than the fluid thread, and the general case where the liquids are of arbitrary viscosity.

Fluid thread highly viscous

For the limiting case where the fluid thread is much more viscous than the external environment, the viscosity of the external environment falls from the growth rate completely. The growth rate thus becomes only a function of the initial radius of thread, the perturbation wavelength, the surface tension of the thread, and the thread viscosity.

Plotting this, one finds that the longest wavelengths are the most unstable. As importantly, one can note that the viscosity of the fluid thread does not influence which wavelengths will be stable. Viscosity only acts to decrease how rapidly a given perturbation will grow or decay with time.

Examples of when this case would apply are when almost any liquid undergoes thread/jet breakup in an air environment.

External fluid highly viscous

For the limiting case where the external environment of the fluid thread is much more viscous than the thread itself, the viscosity of the fluid thread falls from the perturbation growth rate completely. The growth rate thus becomes only a function of the initial radius of the thread, the perturbation wavelength, the surface tension of the thread, the viscosity of the external environment, and the second order Bessel functions of the second kind.

If one were to plot the growth rate as a function of the perturbation wavelength, one would find that the most unstable wavelengths again occur at the longest wavelengths and that the viscosity of the external environment would only act to decrease how rapidly a perturbation would grow or decay in time.

Examples of when this case would apply are when gas bubbles enter a liquid or when water falls into honey.

General case - arbitrary viscosity ratio

The general case for two viscous fluids is much more difficult to solve directly. Tomotika expressed his solution as:

where was defined as:

The coefficients are most easily expressed as the determinants of the following matrices:

The resulting solution remains a function of both the thread and external environment viscosities as well as the perturbation wavelength. The most unstable combination of viscosities and perturbation occurs when with .

For most applications, use of the general case is unnecessary as the two fluids in question have significantly different viscosities which permits the use of one of the limiting cases. However, some instances such as the mixing of oils or oils and water may require the use of the general case.

Satellite drop formation

Water flows from a faucet, producing both a single large droplet and several satellite droplets. Water drop animation enhanced small.gif
Water flows from a faucet, producing both a single large droplet and several satellite droplets.

Satellite drops, also known as secondary droplets, are the drops produced during the thread breakup process in addition to the large main droplet. The drops result when the filament by which the main droplet in hanging from the larger fluid mass itself breaks off from the fluid mass. The fluid contained in the filament can stay as a single mass or breakup due to the recoil disturbances imposed on it by the separation of the main droplet. While the production of satellite droplets can be predicted based on fluid properties, their precise location and volume cannot be predicted. [20] [21]

In general, secondary droplets are an unwanted phenomenon, especially in applications where precise deposition of droplets is important. The production of satellite droplets is governed by the non-linear dynamics of the problem near the final stages of thread breakup.

Examples

The viscosity of honey is large enough to damp all surface perturbations that would lead to the breakup of the thread into droplets. Filtering of honey.jpg
The viscosity of honey is large enough to damp all surface perturbations that would lead to the breakup of the thread into droplets.

Numerous examples of the breakup of fluid threads exist in daily life. It is one of the most common fluid mechanics phenomena one experiences and as such most give the process little thought.

Flow from a faucet

Dripping water is an everyday occurrence. As water leaves the faucet, the filament attached to the faucet begins to neck down, eventually to the point that the main droplet detaches from the surface. [22] The filament cannot retract sufficiently rapidly to the faucet to prevent breakup and thus disintegrates into several small satellite drops. [22]

Air bubbles

Air bubbles are another common breakup phenomenon. As air enters a tank of liquid, like a fish tank, the thread again necks at the base to produce a bubble. Blowing bubbles from a straw into a glass behaves in much the same manner.

Pitch drop experiment

The pitch drop experiment is a famous fluid breakup experiment using high viscous tar pitch. The rate of breakup is slowed to such a degree that only 11 drops have fallen since 1927.

Drops of honey

Honey is sufficiently viscous that the surface perturbations that lead to breakup are almost fully damped from honey threads. This results in the production of long filaments of honey rather than individual droplets.

Related Research Articles

In fluid mechanics, the Grashof number is a dimensionless number which approximates the ratio of the buoyancy to viscous forces acting on a fluid. It frequently arises in the study of situations involving natural convection and is analogous to the Reynolds number.

In fluid dynamics, Stokes' law is an empirical law for the frictional force – also called drag force – exerted on spherical objects with very small Reynolds numbers in a viscous fluid. It was derived by George Gabriel Stokes in 1851 by solving the Stokes flow limit for small Reynolds numbers of the Navier–Stokes equations.

<span class="mw-page-title-main">Boundary layer</span> Layer of fluid in the immediate vicinity of a bounding surface

In physics and fluid mechanics, a boundary layer is the thin layer of fluid in the immediate vicinity of a bounding surface formed by the fluid flowing along the surface. The fluid's interaction with the wall induces a no-slip boundary condition. The flow velocity then monotonically increases above the surface until it returns to the bulk flow velocity. The thin layer consisting of fluid whose velocity has not yet returned to the bulk flow velocity is called the velocity boundary layer.

A Newtonian fluid is a fluid in which the viscous stresses arising from its flow are at every point linearly correlated to the local strain rate — the rate of change of its deformation over time. Stresses are proportional to the rate of change of the fluid's velocity vector.

Hemorheology, also spelled haemorheology, or blood rheology, is the study of flow properties of blood and its elements of plasma and cells. Proper tissue perfusion can occur only when blood's rheological properties are within certain levels. Alterations of these properties play significant roles in disease processes. Blood viscosity is determined by plasma viscosity, hematocrit and mechanical properties of red blood cells. Red blood cells have unique mechanical behavior, which can be discussed under the terms erythrocyte deformability and erythrocyte aggregation. Because of that, blood behaves as a non-Newtonian fluid. As such, the viscosity of blood varies with shear rate. Blood becomes less viscous at high shear rates like those experienced with increased flow such as during exercise or in peak-systole. Therefore, blood is a shear-thinning fluid. Contrarily, blood viscosity increases when shear rate goes down with increased vessel diameters or with low flow, such as downstream from an obstruction or in diastole. Blood viscosity also increases with increases in red cell aggregability.

In fluid dynamics, the capillary number (Ca) is a dimensionless quantity representing the relative effect of viscous drag forces versus surface tension forces acting across an interface between a liquid and a gas, or between two immiscible liquids. Alongside the Bond number, commonly denoted , this term is useful to describe the forces acting on a fluid front in porous or granular media, such as soil. The capillary number is defined as:

Darcy's law is an equation that describes the flow of a fluid through a porous medium. The law was formulated by Henry Darcy based on results of experiments on the flow of water through beds of sand, forming the basis of hydrogeology, a branch of earth sciences. It is analogous to Ohm's law in electrostatics, linearly relating the volume flow rate of the fluid to the hydraulic head difference via the hydraulic conductivity. In fact, the Darcy's law is a special case of the Stokes equation for the momentum flux, in turn deriving from the momentum Navier-Stokes equation.

<span class="mw-page-title-main">Marangoni effect</span> Physical phenomenon between two fluids

The Marangoni effect is the mass transfer along an interface between two phases due to a gradient of the surface tension. In the case of temperature dependence, this phenomenon may be called thermo-capillary convection.

<span class="mw-page-title-main">Weber number</span> Dimensionless number in fluid mechanics

The Weber number (We) is a dimensionless number in fluid mechanics that is often useful in analysing fluid flows where there is an interface between two different fluids, especially for multiphase flows with strongly curved surfaces. It is named after Moritz Weber (1871–1951). It can be thought of as a measure of the relative importance of the fluid's inertia compared to its surface tension. The quantity is useful in analyzing thin film flows and the formation of droplets and bubbles.

Fluid mechanics is the branch of physics concerned with the mechanics of fluids and the forces on them. It has applications in a wide range of disciplines, including mechanical, aerospace, civil, chemical, and biomedical engineering, as well as geophysics, oceanography, meteorology, astrophysics, and biology.

<span class="mw-page-title-main">Multiphase flow</span>

In fluid mechanics, multiphase flow is the simultaneous flow of materials with two or more thermodynamic phases. Virtually all processing technologies from cavitating pumps and turbines to paper-making and the construction of plastics involve some form of multiphase flow. It is also prevalent in many natural phenomena.

The Kelvin equation describes the change in vapour pressure due to a curved liquid–vapor interface, such as the surface of a droplet. The vapor pressure at a convex curved surface is higher than that at a flat surface. The Kelvin equation is dependent upon thermodynamic principles and does not allude to special properties of materials. It is also used for determination of pore size distribution of a porous medium using adsorption porosimetry. The equation is named in honor of William Thomson, also known as Lord Kelvin.

<span class="mw-page-title-main">Plateau–Rayleigh instability</span> Fluid breakup of a falling stream

In fluid dynamics, the Plateau–Rayleigh instability, often just called the Rayleigh instability, explains why and how a falling stream of fluid breaks up into smaller packets with the same volume but less surface area. It is related to the Rayleigh–Taylor instability and is part of a greater branch of fluid dynamics concerned with fluid thread breakup. This fluid instability is exploited in the design of a particular type of ink jet technology whereby a jet of liquid is perturbed into a steady stream of droplets.

<span class="mw-page-title-main">Nonlinear acoustics</span>

Nonlinear acoustics (NLA) is a branch of physics and acoustics dealing with sound waves of sufficiently large amplitudes. Large amplitudes require using full systems of governing equations of fluid dynamics and elasticity. These equations are generally nonlinear, and their traditional linearization is no longer possible. The solutions of these equations show that, due to the effects of nonlinearity, sound waves are being distorted as they travel.

In nonideal fluid dynamics, the Hagen–Poiseuille equation, also known as the Hagen–Poiseuille law, Poiseuille law or Poiseuille equation, is a physical law that gives the pressure drop in an incompressible and Newtonian fluid in laminar flow flowing through a long cylindrical pipe of constant cross section. It can be successfully applied to air flow in lung alveoli, or the flow through a drinking straw or through a hypodermic needle. It was experimentally derived independently by Jean Léonard Marie Poiseuille in 1838 and Gotthilf Heinrich Ludwig Hagen, and published by Hagen in 1839 and then by Poiseuille in 1840–41 and 1846. The theoretical justification of the Poiseuille law was given by George Stokes in 1845.

<span class="mw-page-title-main">Viscosity</span> Resistance of a fluid to shear deformation

The viscosity of a fluid is a measure of its resistance to deformation at a given rate. For liquids, it corresponds to the informal concept of "thickness": for example, syrup has a higher viscosity than water. Viscosity is defined scientifically as a force multiplied by a time divided by an area. Thus its SI units are newton-seconds per square meter, or pascal-seconds.

<span class="mw-page-title-main">Saffman–Taylor instability</span>

The Saffman–Taylor instability, also known as viscous fingering, is the formation of patterns in a morphologically unstable interface between two fluids in a porous medium, described mathematically by Philip Saffman and G. I. Taylor in a paper of 1958. This situation is most often encountered during drainage processes through media such as soils. It occurs when a less viscous fluid is injected, displacing a more viscous fluid; in the inverse situation, with the more viscous displacing the other, the interface is stable and no instability is seen. Essentially the same effect occurs driven by gravity if the interface is horizontal and separates two fluids of different densities, the heavier one being above the other: this is known as the Rayleigh-Taylor instability. In the rectangular configuration the system evolves until a single finger forms, whilst in the radial configuration the pattern grows forming fingers by successive tip-splitting.

The viscous stress tensor is a tensor used in continuum mechanics to model the part of the stress at a point within some material that can be attributed to the strain rate, the rate at which it is deforming around that point.

In fluid mechanics, the thin-film equation is a partial differential equation that approximately predicts the time evolution of the thickness h of a liquid film that lies on a surface. The equation is derived via lubrication theory which is based on the assumption that the length-scales in the surface directions are significantly larger than in the direction normal to the surface. In the non-dimensional form of the Navier-Stokes equation the requirement is that terms of order ε2 and ε2Re are negligible, where ε ≪ 1 is the aspect ratio and Re is the Reynolds number. This significantly simplifies the governing equations. However, lubrication theory, as the name suggests, is typically derived for flow between two solid surfaces, hence the liquid forms a lubricating layer. The thin-film equation holds when there is a single free surface. With two free surfaces, the flow must be treated as a viscous sheet.

Capillary breakup rheometry is an experimental technique used to assess the extensional rheological response of low viscous fluids. Unlike most shear and extensional rheometers, this technique does not involve active stretch or measurement of stress or strain but exploits only surface tension to create a uniaxial extensional flow. Hence, although it is common practice to use the name rheometer, capillary breakup techniques should be better addressed to as indexers.

References

  1. da Vinci, Leonardo (1958). MacCurdy,Edward (ed.). The Notebooks of Leonardo da Vinci. Vol. 2. New York, New York, USA: George Braziller. p.  748.
  2. de Laplace, P.S. (1805). Mechanique Celeste Supplement au X Libre. Paris: Courier.
  3. Young, T (1805). "An Essay on the Cohesion of Fluids". Philosophical Transactions of the Royal Society of London. 95: 65–87. doi: 10.1098/rstl.1805.0005 . S2CID   116124581.
  4. See:
  5. Savart, Félix (1833). "Mémoire sur la constitution des veines liquides lancées par des orifices circulaires en mince paroi" [Memoir on the form of liquid streams issuing from circular orifices in a thin wall]. Annales de chimie et de physique. 2nd series (in French). 53: 337–386.
  6. Plateau, J. (1850). "Ueber die Gränze der Stabilität eines flüssigen Cylinders" [On the limit of stability of a fluid cylinder]. Annalen der Physik und Chemie. 2nd series (in German). 80 (8): 566–569. Bibcode:1850AnP...156..566P. doi:10.1002/andp.18501560808.
  7. Magnus, G. (1859). "Hydraulische Untersuchungen; zweiter Theil" [Hydraulic investigations; second part]. Annalen der Physik und Chemie. 2nd series (in German). 106 (1): 1–32. Bibcode:1859AnP...182....1M. doi:10.1002/andp.18591820102.
  8. Lenard, Philipp (1887). "Ueber die Schwingungen fallender Tropfen" [On the oscillations of falling drops]. Annalen der Physik und Chemie. 3rd series (in German). 30 (2): 209–243. Bibcode:1887AnP...266..209L. doi:10.1002/andp.18872660202.
  9. Dommermuth, DG; Yue DKP (1987). "Numerical simulations of nonlinear axisymmetric flows with a free surface". Journal of Fluid Mechanics. 178: 195–219. Bibcode:1987JFM...178..195D. doi:10.1017/s0022112087001186.
  10. Schulkes, RMS (1994). "The evolution of capillary fountains". Journal of Fluid Mechanics. 261: 223–252. Bibcode:1994JFM...261..223S. doi:10.1017/s0022112094000327.
  11. Youngren, GK; Acrivos A (1975). "Stokes flow past a particle of arbitrary shape: a numerical method of solution". Journal of Fluid Mechanics. 69 (2): 377–403. Bibcode:1975JFM....69..377Y. doi:10.1017/s0022112075001486.
  12. Stone, HA; Leal LG (1989). "Relaxation and breakup of an initially extended drop in an otherwise quiescent fluid" (PDF). Journal of Fluid Mechanics. 198: 399. Bibcode:1989JFM...198..399S. doi:10.1017/s0022112089000194.
  13. Fromm, JE (1984). "Numerical Calculation of the Fluid Dynamics of Drop-on-Demand Jets". IBM Journal of Research and Development. 28 (3): 322–333. doi:10.1147/rd.283.0322.
  14. 1 2 Plateau, J (1850). "Ueber die Gränze der Stabilität eines flüssigen Cylinders". Annalen der Physik. 80 (8): 566–569. Bibcode:1850AnP...156..566P. doi:10.1002/andp.18501560808.
  15. Ting, L; Keller JB (1990). "Slender Jets and Thin Sheets with Surface Tension". SIAM Journal on Applied Mathematics. 50 (6): 1533–1546. doi:10.1137/0150090.
  16. Papageorgiou, DT (1995). "On the breakup of viscous liquid threads". Physics of Fluids. 7 (7): 1529–1544. Bibcode:1995PhFl....7.1529P. CiteSeerX   10.1.1.407.478 . doi:10.1063/1.868540.
  17. Lister, JR; Stone HA (1998). "Capillary breakup of a viscous thread surrounded by another viscous fluid". Physics of Fluids. 10 (11): 2758–2764. Bibcode:1998PhFl...10.2758L. doi:10.1063/1.869799.
  18. Rayleigh, Lord (1892). "XVI. On the instability of a cylinder of viscous liquid under capillary force". Philosophical Magazine. 34 (207): 145–154. doi:10.1080/14786449208620301.
  19. Tomotika, S (1935). "On the Instability of a Cylindrical Thread of a Viscous Liquid Surrounded by Another Viscous Fluid". Proceedings of the Royal Society of London A. 150 (870): 322–337. Bibcode:1935RSPSA.150..322T. doi: 10.1098/rspa.1935.0104 .
  20. Singh, Gaurav (20 January 2013). "Satellite Drop Formation" . Retrieved 18 November 2013.
  21. Henderson, D; Pritchard W; Smolka Linda (1997). "On the pinch-off of a pendant drop of viscous fluid". Physics of Fluids. 9 (11): 3188. Bibcode:1997PhFl....9.3188H. doi:10.1063/1.869435.
  22. 1 2 "Liquid Jet Breakup". The Optical Trek. 2012-12-12. Retrieved 2021-09-29.